Next Article in Journal
The Stability Analysis of Periodic Beams Interacting with Periodic Elastic Foundation with the Use of the Tolerance Averaging Technique
Next Article in Special Issue
A Comprehensive Analysis of the UVC LEDs’ Applications and Decontamination Capability
Previous Article in Journal
Formation of Dislocations and Stacking Faults in Embedded Individual Grains during In Situ Tensile Loading of an Austenitic Stainless Steel
Previous Article in Special Issue
Colored Tattoo Ink Screening Method with Optical Tissue Phantoms and Raman Spectroscopy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Raman Spectroscopic Study of TiO2 Nanoparticles’ Effects on the Hemoglobin State in Individual Red Blood Cells

1
Department of Physics, National Dong Hwa University, Hualien 974301, Taiwan
2
P.N. Lebedev Physical Institute of Russian Academy of Sciences, 119991 Moscow, Russia
3
Department of Physics, Moscow State University, 119991 Moscow, Russia
*
Authors to whom correspondence should be addressed.
Materials 2021, 14(20), 5920; https://doi.org/10.3390/ma14205920
Submission received: 1 September 2021 / Revised: 28 September 2021 / Accepted: 6 October 2021 / Published: 9 October 2021
(This article belongs to the Special Issue Advanced Materials for Biophotonics Applications)

Abstract

:
Titanium dioxide (TiO2) is considered to be a nontoxic material and is widely used in a number of everyday products, such as sunscreen. TiO2 nanoparticles (NP) are also considered as prospective agents for photodynamic therapy and drug delivery. These applications require an understanding of the potential effects of TiO2 on the blood system and its components upon administration. In the presented work, we analyze the interaction of TiO2 nanoparticles of different crystal phases (anatase and rutile) with individual rat Red Blood Cells (RBC) and the TiO2 influence on the oxygenation state and functionality of RBC, estimated via analysis of Raman spectra of Hemoglobin (Hb) and their distribution along individual RBC. Raman spectral signals also allow localization of the TiO2 NP on the RBC. No penetration of the NP inside RBC was observed; however, both kinds of TiO2 NP adsorbed on the RBC membrane can affect the Hb state. Mechanisms involving the NP–membrane–Hb interaction, resulting in partial deoxygenation of Hb and TiO2 photothermal effect on Hb under Raman laser excitation, are suggested. The possible influence on the safety of TiO2 use in advanced medical application, especially on the safety and efficiency of photothermal therapy, is discussed.

1. Introduction

Current perspectives on developing nanoparticles’ (NP) applications for theranostics [1,2] require an understanding of NP interaction with, and potential effects on, the blood system and its components upon administration via injection. NP can also enter the blood stream after administration through breath, dermal deposition, and absorption in the gastrointestinal system. The interaction of various NP with blood and blood components has been studied. The state and functionality of the blood is determined in a significant degree by the state and functionality of red blood cells (RBC). Thus, one of the important criteria for NP use is how the NP can interact with RBC and affect the RBC functionalities. The influence of different kinds of NP on RBC under various conditions was studied, including carbon [3,4,5], silica [6,7], gold and silver [8,9,10,11], Fe3O4 [12] NP, and others [8]; mechanisms of interaction and the conditions of safe application were discussed. Although among other NP, titanium dioxide (TiO2) is often considered as a nontoxic material and is widely used in everyday products of pharmaceutics, cosmetics, and food industries, the problem of its safety when interacting with blood becomes relevant due to current research concerning applications of TiO2 NP for photodynamic therapy and drug delivery [13].
The interaction of TiO2 with RBC has been studied in a number of works [5,8,14,15,16,17,18], and in most of them different kinds of damaging effects were observed, from morphology abnormalities [17], hemolysis [14], membrane properties modification [18], and increasing procoagulant activity of RBC [15], up to induced cell death, genotoxicity, and inflammation [5,19], which depended on internalization pathways, ROS production, and the abnormal interaction of proteins [5]. The effect of TiO2 NP on blood was also observed in in vivo studies [20]. Therefore, TiO2 NP exhibit potential toxicity, and an understanding is needed of how and under which conditions TiO2 can be used to avoid fatal effects or, on the other hand, to induce toxicity for nanomedical purposes.
The protein hemoglobin (Hb) contained in RBC is responsible for the function of blood gas transportation. Violations or modifications of this function by NP can be observed [3,4,10] and need to be controlled. To investigate these effects, noncontact and noninvasive means are essential. For this purpose, Raman spectroscopy is a suitable tool for analysis of the Hb function and its dynamics due to a selective enhancement of the Raman active vibrations in different forms of Hb [3,4,10,21,22,23]. Additionally, Raman mapping allows for a direct observation of the distribution of Hb in different states. It has been applied for analyzing the formation of Hb derivatives in RBC under varying conditions [21,24], studying the process of erythrophagocytosis of senescent RBC [23], and for diagnosing malaria [24].
Raman spectroscopy and Raman mapping were shown to be effective tools for analysis of NP–RBC interaction. Previously, the effects of NP on the oxygenation state of individual RBC were studied for nanodiamond (ND), and it was shown that the adhesion of ND on the RBC membrane affects the cell functioning [3,4,18], but safe conditions for applications can be selected. NDs of different sizes stick to the RBC membrane, slightly affect the deoxygenation degree of RBC, and influence the oxygenation–deoxygenation dynamics of RBC in size-, concentration-, and surface-chemistry-dependent ways [3,4]. More serious disturbance effects were observed in the study of Au and Ag NP interaction with RBC [10]. Increasing hemoglobin deoxygenation was observed in both Au and Ag NP-treated RBC. These studies suggest that the adhesion of NP on the cell membrane causes an imbalance in RBC functioning and cell damage [10,11]. Independent of size and chemical composition, NP penetration into RBC was not observed in these works. In contrast to this, the penetration through the RBC membrane of NP and their aggregates with sizes less than 200 nm was observed using fluorescence microscopy for polymer (with fluorescent dye) NP and via electron microscopy for gold and TiO2 anatase NP after incubation for 4–48 h [8]. Larger particles or aggregates were found stuck to the RBC surface.
In the present work, we analyze the interaction of TiO2 NP of different crystal phases (anatase and rutile) with individual rat RBC. The oxygenation state and functionality of RBC is estimated via analysis of Raman spectra from the Hb Raman mapping of an individual RBC. Raman mapping allows observation of the distribution of Hb in different oxygenation states in the studied RBC. Simultaneously, Raman spectra allow localization of the TiO2 NP on the RBC. In this work, no penetration of the NP inside the RBC was observed; however, both kinds of TiO2 NP were adsorbed on the RBC membrane, which may alter the oxygenation degree in the conditions when RBC should be oxygenated or, in general, could affect the Hb form. The effects of rutile and anatase TiO2 NP are compared. Possible mechanisms of influence of the membrane-localized TiO2 NP on Hb are discussed.

2. Materials and Methods

To estimate the effect of TiO2 nanoparticles on the RBC oxy/deoxy state, two kinds of TiO2 NP with predominant rutile, TiO2(r) (Sigma, Saint Louis, MO, USA), and anatase, TiO2(a), (UV100 Hombikat, Sachtleben Chemie GmbH, Duisburg, Germany) structures were used. The nominal size of TiO2(r) particles was ≤100 nm, and that of TiO2(a) was about 10 nm. In the experiments, the particles’ size and ξ-potential were estimated using the dynamic light scattering method (DLS) with a Zetasizer Nano ZS (Malvern Instruments, Malvern, UK). NP were suspended in ultrapure MilliQ water (Type 1, Direct-Q 3UV, Merck-Millipore, Burlington, MA, USA) at pH 6 and in PBS (pH 7.3) in concentration 0.1 mg/mL. Absorption spectra were measured with a UV–visible spectrometer to confirm high absorption from UV up to 450 nm range. Structural phases of TiO2 NP were confirmed by Raman spectra measurements with an α-SNOM Raman microspectrometer (Witec, Ulm, Germany), with excitation by an Ar ion laser with 488 nm wavelength.
For the rat red blood cell (RBC) samples, 3 mL of whole blood was withdrawn from the tails of Wistar rats and transferred into EDTA-covered tubes. The research methods were approved by the Animal Care and Use Committee of National Dong Hwa University (Approval ID100004). The rats were narcotized for the experiments. RBC samples were prepared and treated with NP as described previously [3,4]. In short, the RBC were separated from the fresh whole blood using centrifugation, repeatedly washed with standard phosphate buffer saline (pH 7.4), then RBC mass was diluted with PBS at the ratio of 5:1000 µL (RBC:PBS). The spectra of RBC in an oxygenated state were measured in ambient conditions and after nitrogen gas purging for deoxygenated RBC.
The TiO2 powders were diluted in PBS at 33 mg/mL concentration. The obtained suspension was added to the RBC:PBS sample and coincubated for 1 h with the final concentration of TiO2 ~20 µg/mL in the prepared sample. The prepared sample suspensions of RBC with TiO2 powders and control RBC without TiO2 were placed on a Si substrate for Raman measurements. The samples were at ambient atmospheric pressure, so RBC were oxygenated [3,4]. Raman spectra were measured and the selected Raman signal intensity mapping was performed with α-SNOM Raman microspectrometer (Witec, Ulm, Germany), with excitation by an Ar ion laser of 488 nm wavelength, objective ×50, power ~0.3 mW at the output, scanning time 150–220 s in ambient atmosphere conditions. Raman mapping of 15–20 RBCs from each sample were performed. Every mapping contained 24 × 24 spectra which were analyzed using software of Witec α-SNOM ScanCntr Spectroscopy Plus 1.34 (Witec, Ulm, Germany) and then with Origin 9.1 (OriginLab, Northampton, MA, USA).

3. Results

3.1. TiO2 Particles Characterization

The size and ζ-potential of rutile TiO2(r) and anatase TiO2(a) phases NP suspended in ultrapure MilliQ water were measured. The ζ-potentials were found to be 15.5 ± 0.37 mV for anatase NP and 11.1 ± 0.36 mV for rutile. The size of TiO2(r) at that condition showed a narrow distribution with a maximum near 154.4 ± 16.1 nm; TiO2(a) revealed two fractions, one centered near size 136 nm and the other, characterizing the aggregates, with an average size of 448.1 ± 27.3 nm. When dissolved in PBS (pH 7.3), the measured values of ζ-potentials became negative and equal to −23.4 ± 2.0 and −25.2 ± 1.43 mV, respectively for the TiO2(r) and anatase TiO2(a). From the particle size distribution, the TiO2 in PBS aggregated or agglomerated and formed a wide distribution with the hydrodynamic size up to one micron (data not shown); however, some crucially decreased fraction of small particles or small aggregates still exists. We deliberately did not remove the larger aggregates, as we wanted to examine the more practical application when RBC encounter all kinds of TiO2 aggregates, especially in PBS. The particles’ agglomeration significantly increased in PBS in comparison with water, as was shown for TiO2 NP in solutions with high ionic strength at pH 7.3 [25]. In such studies, the primary size of the TiO2 NP, usually tens of nm, together with the crystallinity, played a role in the properties of agglomerates, and the properties inherent in nanoscale particles could also manifest at TiO2 interaction with cells. The observed average hydrodynamic size of agglomerates reached 751.2 ± 48.8 nm (rutile) and 673 ± 29.0 nm (anatase). The agglomeration and changes in ζ-potential may have happened due to the influence of PBS components, as has been shown previously [26], where ions such as Cl, Na+, and PO4−3 from PBS can affect the particle surface properties and stabilization of TiO2 NP suspension.
Typical Raman spectra of the studied TiO2 powders are shown in Figure 1. Characteristic peaks were observed and agreed with literature values [27]. The spectra confirm the structure of two kinds of TiO2 NP used in this work. These observed characteristic peaks were used as markers to localize TiO2 in the Raman mappings of RBC at their interaction with TiO2.

3.2. Raman Analysis of the RBC Oxygenation State

The red blood cells’ main component is Hb, the iron-containing oxygen transporting protein. Raman spectroscopy is used for studies of Hb and RBC; aside from its low-invasive nature, it allows for the simultaneous identification of oxygenated and deoxygenated states of Hb. Hemoglobin’s Raman spectrum is strong and complex but with fingerprints from heme (the active center of Hb) and protein units [21,22,24]. The spectra of oxygenated and deoxygenated RBC are shown in Figure 2. These spectra were measured from individual RBC with 488 nm wavelength laser excitation. Raman vibrations in different oxygenation states exhibited obvious changes in the oxidation state sensitive region (1350 to 1380 cm−1, ν4 band), spin state sensitive region (1500 to 1650 cm−1, ν19, ν37, and ν10 bands), and the C-H deformation region (1200 to 1250 cm−1), followed by T ↔ R (tense–relaxed) transition in the Hb molecule in the oxygenation–deoxygenation process [22,28,29]. Based on the light absorption spectrum of oxyHb and deoxyHb, excitation can be selected close to the absorption bands to result in selective enhancement of resonance Raman peaks [30]. The Soret band maxima of oxy- and deoxy-Hb were 415 and 430 nm, respectively, and it has been shown that at 488 nm wavelength laser excitation, clear spectral changes in the 1350–1370 cm−1 (ν4 band) can be observed [28]. This band was assigned to the symmetrical pyrrole half-ring stretching vibration and the shift from 1376 cm−1 to 1358 cm−1 was demonstrated as a marker of deoxygenation at resonance Raman study of the oxygenation–deoxygenation process [30]; the peak’s intensity ratio was proportional to the portion of oxygenated Hb [4,30].
Typical spectra for oxyHb and deoxyHb are displayed in Figure 2, with the markers for the Raman mapping. The Raman mapping of RBC treated with TiO2 was performed and compared with the control of the untreated RBC. Figure 3 shows the typical mapping of control RBC (Figure 3a) and RBC treated with TiO2(r) (Figure 3b,c). Raman mapping illustrated the spatial distribution of Raman signal intensity along the sample; the distribution of oxyHb was mapped via Raman peak with a maximum in the 1365–1380 cm−1 range (O in Figure 2) and distribution of deoxyHb via a peak in the 1350–1360 cm−1 range (D in Figure 2). TiO2 was visualized via mapping of Raman signal in the 600–650 cm-1 range where A1g band of Rutile (610 cm−1) and Eg band of Anatase (639 cm-1) were revealing, peaks shown in Figure 1. Note that the measurements were performed in ambient conditions; thus, the RBC should be mostly oxygenated. The spectra measured from the TiO2-treated and untreated RBC allowed us to observe and discuss the effect of TiO2 NP on oxygenated RBC, while mapping of the intensity of Raman signal in characteristic RBC (Hb) and TiO2 ranges allowed, correspondingly, visualizing the relative distributions of oxygenated oxyHb (Figure 3-I), deoxygenated deoxyHb (Figure 3-II), and TiO2 localization at interaction with the RBC (Figure 3b,c-III) in comparison with the control (Figure 3a-III). Note that TiO2 is not fluorescent and cannot be easily observed with the methods of laser fluorescence scanning microscopy, while Raman mapping serves as an alternative means of TiO2 visualization.
In the images of a control cell (Figure 3a) the part of oxygenated Hb was localized in the center of the RBC (Figure 3a-I), the relatively deoxygenated Hb was distributed on the peripheral part of RBC (Figure 3a-II). In general, a similar distribution was observed for mapping of RBCs treated with TiO2. In Figure 3b,c, Raman mappings of an RBC treated with TiO2(r) and TiO2(a) are presented, and the maps of relatively oxygenated Hb (Figure 3b,c-I) and relatively deoxygenated Hb (Figure 3b,c-II) were added with a distribution of TiO2 (Figure 3b,c-III). Well-detectable aggregates of TiO2 localized on the outer surface of the RBC membrane were observed in both cases of using Anatase and Rutile particles. In spite of a strong Raman signal of TiO2, we did not observe any NP which could be considered localized inside the cell.
Both Rutile and Anatase NP were tested before, and the toxicity of Rutile and Anatase could differ. In different studies, the results were controversial and depended significantly on the experimental conditions [31,32,33,34]. In order to discuss the possible effect of both kinds of TiO2 NP on RBC, the spectra were analyzed focusing on the areas of maximum oxygenation (as shown in Figure 3-I) and maximum deoxygenation (as in Figure 3-II) in the Raman maps of each studied individual RBC. In Figure 4, the spectra of individual RBCs in a narrow range containing the band ν4 are shown. A large number of individual cells was analyzed, the variability in Hb oxygenation state distribution along one cell and an average oxygenation degree for the different cells was observed. The spectra for each analyzed RBC were selected to demonstrate the highest (I) and lowest (II) oxygenation degree, which could be observed in the same individual cell (Figure 4a). The spectra in Figure 4a,b were measured from similar areas of two different cells and show the oxygenation degree of variability between cells. Figure 4(1) shows spectra for control untreated RBCs, while Figure 4(2,3) show spectra for RBCs treated with Rutile TiO2(r) and TiO2(a), respectively.
Figure 4 shows the data for cells indicating the largest difference found in the oxygenation degree inside one cell differs for untreated and TiO2-treated RBC. The spectra shown for higher (I) and lower (II) oxygenation states of control RBC (Figure 4(1)) with the maximum at 1372 cm−1 indicated mostly oxygenated Hb. Only a weakly expressed shoulder (arising from the peak at 1356 cm−1) was observed in the spectra (II) both in Figure 4a(1),b(1), so the content of deoxygenated Hb can be considered very low. However, this low content of particularly deoxygenated Hb was still observable with Raman mapping in Figure 3a-II.
As the Raman peak intensity directly depends on the number of oscillators, the oxygenation degree SO2 can be considered proportional to the ratio of intensities I1372 and I1356 [4,30], and can be expressed as SO2 = A × I1372/(I1372 + I1356) + B, where I is the peak intensity at the corresponding wavenumber, and A and B are coefficients, which have to be determined via a calibration with independent oximetry measurements [28].
Figure 4a(1),b(1) show that the control RBC were predominantly oxygenated, as expected in ambient conditions. I1372/(I1372 + I1356) can be considered close to 1 in the RBC center (as in Figure 3a), and it decreased in the thin layer on the periphery in the cell (Figure 3b). Figure 4(2,3) show the spectra for RBC treated with different kinds of TiO2. In contrast to Figure 4(1), in Figure 4a(2),a(3), oxyHb in the center of the cell, with a ratio I1372/(I1372+ I1356) close to 1, coexisted with significantly deoxygenated Hb on the cell periphery. An even lower oxygenation degree was estimated along the whole cells, presented by spectra in Figure 4b(2),b(3), where the central part of RBC contained only particularly oxygenated Hb, with both well-observable peaks at 1372 and 1356 cm−1, while the periphery revealed mostly deoxyHb. Thus, the content of deoxyHb was observed to be higher in both TiO2-treated cells than in the control. Spectra extended in the spin-sensitive range (Figure 5) also confirmed a decrease in oxyHb markers (ν37 at 1588 cm−1 and ν10 at 1640 cm1) for TiO2-treated RBC (Figure 5b,c) in comparison with untreated cells (Figure 5a). This means that TiO2 adsorbed on the RBC membrane may affect Hb oxygenation degree in the air environment when RBC should be oxygenated. Analyzing the variability of TiO2 effect on oxygenated RBC, no difference between the Rutile and Anatase particles was observed. Maxima for ν4 in oxygenated and deoxygenated states were determined via the spectra deconvolution (Figure 6), and the value I1372/(I1372 + I1356) characterizing the variability of RBC oxygenation for untreated and treated RBC is presented in the inset of Figure 6.

4. Discussion

In previous studies, a shift of the ν4 band in the process of Hb oxygenation was observed [30], and it was demonstrated that the ratio of intensities of the peaks at 1372 cm−1 and 1356 cm−1 could be used as a marker of the oxygenation state of Hb [10,21,30]. Additionally, in Raman mapping, we observed the inhomogeneous distribution of Hb with various oxygenation degree values along an individual RBC, and analyzed how the adsorption of TiO2 on the RBC membrane affected this distribution. The predominant localization of oxygenated Hb is in the RBC center (Figure 3a-I), while a small fraction of partially deoxygenated Hb is distributed on the peripheral part of the control RBC (Figure 3a-II). It is consistent with the studies of deoxyHb binding on RBC membrane as a result of electrostatic interactions, covalent association with the membrane components by disulfide bonds, and adsorption to membrane lipids via hydrophobic interactions [35]. In particular, the reversible association of deoxyHb with the RBC membrane via interaction with band 3 protein provides an active role of the membrane in RBC gases’ transport function [35]. It is considered a “molecular switch” mechanism for regulating RBC biology (metabolic and transport processes, signal transduction, etc.) [36]. Similarly, RBCs treated with both kinds of TiO2 demonstrated the peripheral localization of some portion of deoxyHb and the central localization of Hb with a higher oxygenation degree (Figure 3b, for TiO2(r) and Figure 3c for TiO2(a)). However, the oxygenation degree of the whole TiO2-treated RBC was decreased as is shown in the inset of Figure 6.
As the plasma membrane of RBC plays a fundamental role in the oxygen (and other gases) transport by RBC and in maintaining RBC metabolism [37], the membrane structural changes, which can be created by adsorbed NP aggregates, can affect the RBC state [3,4,10,11,18]. Observed TiO2 aggregates or agglomerates on the RBC membrane agreed, e.g., a previous study showed TiO2 particles sticking to RBC membrane regardless of their surface charge [18]. It suggests the mostly nonelecrostatic nature of the NP–membrane adhesion interaction; however, it is not clear, and more detailed studies are needed to confirm. The membrane permeability for oxygen is determined by the membrane molecules’ mobility; thus, the interaction between the membrane and adhered NP aggregates or agglomerates can affect the conditions for the gas molecules’ transmembrane diffusion and for Hb adsorption on the membrane. Additionally, as was shown [10], the interaction with adsorbed NP may create an obstacle for the exchange of ions and polar molecules via protein channels and also due to production of reactive oxygen species [14] and lipid peroxidation. These reasons can lead to altering cell membrane properties, pH imbalances inside the cells, and finally can affect the oxygen binding affinity of Hb. As a result, similar deoxygenation under NP treatment has been observed [10] for RBC trapped with optical tweezers and treated with Ag or Au NP.
Additionally, using an excitation laser with 488 nm wavelength, we needed to consider the probability of laser effects directly on Hb [38], such as photodegradation of Hb (denaturation, hemichrome formation, and potential aggregation) in the process of Raman measurements. Some researchers suggest that the position of the ν4 peak can indicate the iron ion transition to the ferric form [39], and this is considered as one of the markers of the Hb photodamage [40]. This effect is pronounced in the presence of both kinds of TiO2 in comparison with control. Thus, this combined effect is facilitated by the oxidative stress due to photocatalytic and phototoxic properties of TiO2 at the excitation close to the TiO2 absorption range [41]. Most of the resulting Hb forms contained the ferric heme iron and could hardly be distinguished from oxyHb. However, comparison of the spectra in Figure 4 and Figure 5 shows that the position of the ν4 peak at 1372 cm−1 correlated with the presence of other peaks characterizing oxyHb (ν37 1588 cm−1 and ν10 1640 cm−1), while ν4 observed at 1356 cm−1 correlated with their decreasing. This decrease in intensity of ν37 and ν10 peaks allows us to conclude that particular deoxygenation of Hb takes place, rather than Hb photodegradation. We suggest TiO2 affects Hb state through affecting the membrane properties and stimulats Hb deoxygenation.. We cannot exclude also the input of photodissociation of Hb, also resulting in the observation of higher content of deoxyHb. We suggest also that, together with influence on Hb through affecting the membrane properties and stimulating Hb deoxygenation membrane-associated TiO2, aggregates/agglomerates can promote photo or photothermal dissociation of the oxyHb.
Use of TiO2 NP for nanotherapy is currently widely discussed [13]. The limitations of their application as photosensitizers for photodynamic therapy are discussed in terms of possible side-effects, but they are considered nontoxic for their use in drug delivery or for ultrasound activation of ROS formation for anticancer treatment [42,43]. However, significant and various detrimental influences of TiO2 NP on blood and blood components are demonstrated, and understanding these effects and studying these mechanisms, including the nonphotoactivated ones, is necessary.
In conclusion, the TiO2 effect on RBC was observed, resulting in a decrease in the oxygenation degree of RBC. We suggest that TiO2 affects Hb oxygenation state via altering RBC membrane properties. The disturbing treatment of the laser in the presence of TiO2 is also probable, resulting in oxygen photodissociation of Hb. These effects should be considered in the development of the method of TiO2 NP theranostic applications which imply NP interaction with blood.

Author Contributions

Conceptualization, C.-L.C., E.P., Y.-C.L. and A.P.; methodology, E.P. and Y.-C.L.; analysis, E.P., A.K. and A.L.; investigation, Y.-C.L., K.-T.W. and A.K.; data curation, E.P.; writing E.P.; editing, E.P., A.K., E.S., A.P. and C.-L.C.; visualization, E.P., Y.-C.L., K.-T.W.; supervision, C.-L.C. and A.P.; project administration, C.-L.C. and A.P.; and funding acquisition, C.-L.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Taiwan Ministry of Science and Technology grants MOST 106-2112-M-259-009-MY3, MOST 109-2923-M-259-001-MY3 and the Russian Science foundation grant # 20-45-08004.

Institutional Review Board Statement

The study was conducted according to the guidelines of the Declaration of Helsinki and approved by the Animal Care and Use Committee of National Dong Hwa University (Approval ID100004).

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Urbán, P.; Liptrott, N.J.; Bremer, S. Overview of the blood compatibility of nanomedicines: A trend analysis of in vitro and in vivo studies. WIREs Nanomed. Nanobiotechnol. 2019, 11, e1546. [Google Scholar] [CrossRef]
  2. Zhang, N.; Wei, M.Y.; Ma, Q. Nanomedicines: A Potential Treatment for Blood Disorder Diseases. Front. Bioeng. Biotechnol. 2019, 7, 369. [Google Scholar] [CrossRef] [Green Version]
  3. Lin, Y.C.; Tsai, L.W.; Perevedentseva, E.; Chang, H.H.; Lin, C.H.; Sun, D.S.; Lugovtsov, A.; Priezzhev, A.; Jani, M.; Cheng, C.L. The influence of nanodiamond on the oxygenation states and micro rheological properties of human red blood cells in vitro. J. Biomed. Opt. 2012, 17, 101512. [Google Scholar] [CrossRef] [Green Version]
  4. Tsai, L.W.; Lin, Y.C.; Perevedentseva, E.; Lugovtsov, A.; Priezzhev, A.; Cheng, C.L. Nanodiamonds for Medical Applications: Interaction with Blood in Vitro and in Vivo. Int. J. Mol. Sci. 2016, 17, 1111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Boland, S.; Hussain, S.; Baeza-Squiban, A. Carbon Black and Titanium Dioxide Nanoparticles Induce Distinct Molecular Mechanisms of Toxicity. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2014, 6, 641–652. [Google Scholar] [CrossRef] [Green Version]
  6. Zhao, Y.; Sun, X.; Zhang, G.; Trewyn, B.G.; Slowing, I.I.; Lin, V.S.Y. Interaction of Mesoporous Silica Nanoparticles with Human Red Blood Cell Membranes: Size and Surface Effects. ACS Nano 2011, 5, 1366–1375. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Jiang, L.; Yu, Y.; Li, Y.; Yu, Y.; Duan, J.; Zou, Y.; Li, Q.; Sun, Z. Oxidative Damage and Energy Metabolism Disorder Contribute to the Hemolytic Effect of Amorphous Silica Nanoparticles. Nanoscale Res. Lett. 2016, 11, 57. [Google Scholar] [CrossRef] [Green Version]
  8. Rothen-Rutishauser, B.M.; Schürch, S.; Haenni, B.; Kapp, N.; Gehr, P. Interaction of Fine Particles and Nanoparticles with Red Blood Cells Visualized with Advanced Microscopic Techniques. Environ. Sci. Technol. 2006, 40, 4353–4359. [Google Scholar] [CrossRef] [PubMed]
  9. Purohit, R.; Vallabani, N.V.S.; Shukla, R.K.; Kumar, A.; Singh, S. Effect of gold nanoparticle size and surface coating on human red blood cells. Bioinspired Biomim. Nanobiomater. 2016, 5, 121–131. [Google Scholar] [CrossRef]
  10. Barkur, S.; Lukose, J.; Chidangil, S. Probing Nanoparticle–Cell Interaction Using Micro-Raman Spectroscopy: Silver and Gold Nanoparticle-Induced Stress Effects on Optically Trapped Live Red Blood Cells. ACS Omega 2020, 5, 1439–1447. [Google Scholar] [CrossRef]
  11. Kwon, T.W.; Woo, H.J.; Kim, Y.H.; Lee, H.J.; Park, K.H.; Park, S.; Youn, B.H. Optimizing Hemocompatibility of Surfactant-Coated Silver Nanoparticles in Human Erythrocytes. J. Nanosci. Nanotechnol. 2012, 12, 6168–6175. [Google Scholar] [CrossRef]
  12. Al-Akhras, M.A.H.; Aljarrah, K.; Albiss, B.; Al-Khalili, D. Influence of iron oxide nanoparticles (Fe3O4) on erythrocyte photohemolysis via photofrin and Rose Bengal sensitization. Photodiagnosis Photodyn Ther. 2017, 18, 111–118. [Google Scholar] [CrossRef]
  13. Ziental, D.; Czarczynska-Goslinska, B.; Mlynarczyk, D.T.; Glowacka-Sobotta, A.; Stanisz, B.; Goslinski, T.; Sobotta, L. Titanium Dioxide Nanoparticles: Prospects and Applications in Medicine. Nanomaterials 2020, 10, 387. [Google Scholar] [CrossRef] [Green Version]
  14. Li, S.Q.; Zhu, R.R.; Zhu, H.; Xue, M.; Sun, X.Y.; Yao, S.D.; Wang, S.L. Nanotoxicity of TiO2 nanoparticles to erythrocyte in vitro. Food Chem. Toxicol. 2008, 46, 3626–3631. [Google Scholar] [CrossRef] [PubMed]
  15. Bian, Y.; Chung, H.Y.; Bae, O.N.; Lim, K.M.; Chung, J.H.; Pi, J. Titanium dioxide nanoparticles enhance thrombosis through triggering the phosphatidylserine exposure and procoagulant activation of red blood cells. Part. Fibre Toxicol. 2021, 18, 28. [Google Scholar] [CrossRef] [PubMed]
  16. Liu, K.; Lin, X.; Zhao, J. Toxic effects of the interaction of titanium dioxide nanoparticles with chemicals or physical factors. Int. J. Nanomed. 2013, 8, 2509–2520. [Google Scholar]
  17. Tsui, S.M.; Ahmed, R.; Amjad, N.; Ahmed, I.; Yang, J.; Manno, F.; Barman, I.; Shih, W.C.; Laua, C. Single red blood cell analysis reveals elevated hemoglobin in poikilocytes. J. Biomed. Opt. 2020, 25, 015004. [Google Scholar] [CrossRef] [PubMed]
  18. Avsievich, T.; Popov, A.; Bykov, A.; Meglinski, I. Mutual interaction of red blood cells influenced by nanoparticles. Sci. Rep. 2019, 9, 5147. [Google Scholar] [CrossRef] [PubMed]
  19. Ghosh, M.; Chakraborty, A.; Mukherjee, A. Cytotoxic, genotoxic and the hemolytic effect of titanium dioxide (TiO2) nanoparticles on human erythrocyte and lymphocyte cells in vitro. Appl. Toxicol. 2013, 33, 1097–1110. [Google Scholar] [CrossRef] [PubMed]
  20. Hadei, M.; Rabbani, S.; Nabizadeh, R.; Mahvi, A.H.; Mesdaghinia, A.; Naddafi, K. Comparison of the Toxic Effects of Pristine and Photocatalytically Used TiO2 Nanoparticles in Mice. Biol. Trace. Elem. Res. 2021. [Google Scholar] [CrossRef] [PubMed]
  21. Wood, B.R.; Kochan, K.; Marzec, K.M. Resonance Raman spectroscopy of hemoglobin in red blood cells. In Vibrational Spectroscopy in Protein Research; Ozaki, Y., Baranska, M., Lednev, I., Wood, B., Eds.; Academic Press: Cambridge, UK, 2020; pp. 375–414. [Google Scholar]
  22. Atkins, C.G.; Buckley, K.; Blades, M.W.; Turner, R.F.B. Raman Spectroscopy of Blood and Blood Components. Appl. Spectrosc. 2017, 71, 767–793. [Google Scholar] [CrossRef]
  23. Dybas, J.; Chiura, T.; Marzec, K.M.; Mak, P.J. Probing Heme Active Sites of Hemoglobin in Functional Red Blood Cells Using Resonance Raman Spectroscopy. J. Phys. Chem. B 2021, 125, 3556–3565. [Google Scholar] [CrossRef] [PubMed]
  24. Wood, B.R.; Stoddart, P.R.; McNaughton, D. Molecular Imaging of Red Blood Cells by Raman Spectroscopy. Aust. J. Chem. 2011, 64, 593–599. [Google Scholar] [CrossRef]
  25. Jiang, J.; Oberdoerster, G.; Biswas, P. Characterization of size, surface charge, and agglomeration state of nanoparticle dispersions for toxicological studies. J. Nanopart. Res. 2009, 11, 77–89. [Google Scholar] [CrossRef]
  26. Zhang, C.; Lohwacharin, J.; Takizawa, S. Properties of residual titanium dioxide nanoparticles after extended periods of mixing and settling in synthetic and natural waters. Sci. Rep. 2017, 7, 9943. [Google Scholar] [CrossRef] [Green Version]
  27. Shaikh, S.F.; Mane, R.S.; Min, B.K.; Hwang, Y.J.; Joob, O. D-sorbitol-induced phase control of TiO2 nanoparticles and its application for dye-sensitized solar cells. Sci. Rep. 2016, 6, 20103. [Google Scholar] [CrossRef]
  28. Torres Filho, I.P.; Terner, J.; Pittman, R.N.; Proffitt, E.; Ward, K.R. Measurement of hemoglobin oxygen saturation using Raman microspectroscopy and 532-nm excitation. J. Appl. Physiol. 2008, 104, 1809–1817. [Google Scholar] [CrossRef] [Green Version]
  29. Wood, B.R.; Tait, B.; McNaughton, D. Micro-Raman characterisation of the R to T state transition of haemoglobin within a single living erythrocyte. Biochim. Biophys. Acta Mol. Cell Res. 2001, 1539, 58–70. [Google Scholar] [CrossRef] [Green Version]
  30. Torres Filho, I.P.; Terner, J.; Pittman, R.N.; Somera, L.G.; Ward, K.R. Hemoglobin oxygen saturation measurements using resonance Raman intravital microscopy. Am. J. Physiol. Heart. Circ. Physiol. 2005, 289, H488–H495. [Google Scholar] [CrossRef] [Green Version]
  31. Iavicoli, I.; Leso, V.; Fontana, L.; Bergamachi, A. Toxicological effects of titanium dioxide nanoparticles: A review of in vitro mammalian studies. Eur. Rev. Med. Pharm. Sci. 2011, 15, 481–508. [Google Scholar]
  32. Shah, S.N.A.; Shah, Z.; Hussain, M.; Khan, M. Hazardous effects of titanium dioxide nanoparticles in ecosystem. Bioinorg. Chem. Appl. 2017, 2017, 4101735. [Google Scholar] [CrossRef] [Green Version]
  33. De Matteis, V.; Cascione, M.; Brunetti, V.; Toma, C.C.; Rinaldi, R. Toxicity assessment of anatase and rutile titanium dioxide nanoparticles: The role of degradation in different pH conditions and light exposure. Toxicol. Vitr. 2016, 37, 201–210. [Google Scholar] [CrossRef]
  34. Yu, Q.; Wangm, H.; Peng, Q.; Li, Y.; Liu, Z.; Lia, M. Different toxicity of anatase and rutile TiO2 nanoparticles on macrophages: Involvement of difference in affinity to proteins and phospholipids. J. Hazard. Mater. 2017, 335, 125–134. [Google Scholar] [CrossRef]
  35. Kosmachevskaya, O.V.; Nasybullina, E.I.; Blindar, V.N.; Topunov, A.F. Binding of Erythrocyte Hemoglobin to the Membrane to Realize Signal-Regulatory Function (Review). Appl. Biochem. Microbiol. 2019, 55, 83–98. [Google Scholar] [CrossRef]
  36. Chu, H.; McKenna, M.M.; Krump, N.A.; Zheng, S.; Mendelsohn, L.; Thein, S.L.; Garrett, L.J.; Bodine, D.M.; Low, P.S. Reversible binding of hemoglobin to band 3 constitutes the molecular switch that mediates O2 regulation of erythrocyte properties. Blood 2016, 128, 2708–2716. [Google Scholar] [CrossRef] [PubMed]
  37. De Rosa, M.C.; Carelli Alinovi, C.; Galtieri, A.; Scatena, R.; Giardina, B. The plasma membrane of erythrocytes plays a fundamental role in the transport of oxygen, carbon dioxide and nitric oxide and in the maintenance of the reduced state of the heme iron. Gene 2007, 398, 162–171. [Google Scholar] [CrossRef]
  38. Ahlawat, S.; Kumar, N.; Uppal, A.; Gupta, P.K. Visible Raman excitation laser induced power and exposure dependent effects in red blood cells. J. Biophotonics 2016, 10, 415–422. [Google Scholar] [CrossRef]
  39. Marzec, K.M.; Rygula, A.; Wood, B.R.; Chlopickia, S.; Baranska, M. High-resolution Raman imaging reveals spatial location of heme oxidation sites in single red blood cells of dried smears. J. Raman Spectrosc. 2015, 46, 76–83. [Google Scholar] [CrossRef]
  40. Menżyk, A.; Damin, A.; Martyna, A.; Alladio, E.; Vincenti, M.; Martra, G.; Zador, G. Toward a novel framework for bloodstains dating by Raman spectroscopy: How to avoid sample photodamage and subsampling errors. Talanta 2020, 209, 120565. [Google Scholar] [CrossRef] [PubMed]
  41. Dalai, S.; Pakrashi, S.; Suresh Kumar, R.S.; Chandrasekaran, N.; Mukherjee, A. A comparative cytotoxicity study of TiO2 nanoparticles under light and dark conditions at low exposure concentrations. Toxicol. Res. 2012, 1, 116–130. [Google Scholar] [CrossRef]
  42. You, D.G.; Deepagan, V.G.; Um, W.; Jeon, S.; Son, S.; Chang, H.; Yoon, H.I.; Cho, Y.W.; Swierczewska, M.; Lee, S.; et al. ROS-generating TiO2 nanoparticles for non-invasive sonodynamic therapy of cancer. Sci. Rep. 2016, 6, 23200. [Google Scholar] [CrossRef] [PubMed]
  43. Ou, G.; Li, Z.; Li, D.; Cheng, L.; Liu, Z.; Wu, H. Photothermal therapy by using titanium oxide nanoparticles. Nano Res. 2016, 9, 1236–1243. [Google Scholar] [CrossRef]
Figure 1. Raman spectra of Rutile TiO2(r) and Anatase TiO2(a) NP with 488 nm wavelength laser excitation. The peaks, positions, and modes are marked according to literature values.
Figure 1. Raman spectra of Rutile TiO2(r) and Anatase TiO2(a) NP with 488 nm wavelength laser excitation. The peaks, positions, and modes are marked according to literature values.
Materials 14 05920 g001
Figure 2. Raman spectra of oxyHb and deoxyHb, measured in oxygenated and deoxygenated RBC with 488 nm wavelength laser excitation. (O) and (D) mark the wavenumber ranges for mapping of ν4 band: oxyHb (1365–1380 cm−1) and deoxyHb (1350–1360 cm−1), respectively. Peaks for oxyHb 1588 cm−1 (ν37) and 1640 cm−1 (ν10) are also shown. Peak assignment according to [21].
Figure 2. Raman spectra of oxyHb and deoxyHb, measured in oxygenated and deoxygenated RBC with 488 nm wavelength laser excitation. (O) and (D) mark the wavenumber ranges for mapping of ν4 band: oxyHb (1365–1380 cm−1) and deoxyHb (1350–1360 cm−1), respectively. Peaks for oxyHb 1588 cm−1 (ν37) and 1640 cm−1 (ν10) are also shown. Peak assignment according to [21].
Materials 14 05920 g002
Figure 3. Raman mapping of individual RBC: (a) control RBC; (b) RBC with TiO2 (Rutile); and (c) RBC with TiO2 (Anatase). (I) mapping in the range of 1365–1380 cm−1 shows distribution of oxyHb; (II) mapping in the range 1350–1360 cm−1 shows the distribution of deoxyHb; and (III) mapping in the range 600–650 cm−1 shows the distribution of the TiO2 peaks only in the TiO2-treated RBC. The spectra were collected at ambient conditions, 488 nm wavelength laser excitation; 24 × 24 pixels. Scale bar is 1 μm. The RBC contour in mapping of TiO2 in (III) is marked according to the cell mapping in (I).
Figure 3. Raman mapping of individual RBC: (a) control RBC; (b) RBC with TiO2 (Rutile); and (c) RBC with TiO2 (Anatase). (I) mapping in the range of 1365–1380 cm−1 shows distribution of oxyHb; (II) mapping in the range 1350–1360 cm−1 shows the distribution of deoxyHb; and (III) mapping in the range 600–650 cm−1 shows the distribution of the TiO2 peaks only in the TiO2-treated RBC. The spectra were collected at ambient conditions, 488 nm wavelength laser excitation; 24 × 24 pixels. Scale bar is 1 μm. The RBC contour in mapping of TiO2 in (III) is marked according to the cell mapping in (I).
Materials 14 05920 g003
Figure 4. Characteristic Raman spectra of individual RBCs extended in the 1300–1425 cm−1 range. Examples for (a) the most oxygenated RBCs; and (b) the most deoxygenated RBCs. For every RBC, the spectra demonstrate the higher (I) and lower (II) oxygenation states. (1) Control RBC; (2) RBC treated with TiO2(r); and (3) RBC treated with TiO2(a). Blue and green arrows mark the ν4 band at 1356 cm−1 and 1372 cm−1, respectively.
Figure 4. Characteristic Raman spectra of individual RBCs extended in the 1300–1425 cm−1 range. Examples for (a) the most oxygenated RBCs; and (b) the most deoxygenated RBCs. For every RBC, the spectra demonstrate the higher (I) and lower (II) oxygenation states. (1) Control RBC; (2) RBC treated with TiO2(r); and (3) RBC treated with TiO2(a). Blue and green arrows mark the ν4 band at 1356 cm−1 and 1372 cm−1, respectively.
Materials 14 05920 g004
Figure 5. Characteristic Raman spectra measured in the most oxygenated and in the most deoxygenated areas of an individual RBC: (a) control; (b) RBC treated with TiO2(r); and (c) RBC treated with TiO2(a). Spectra are extended in the 1500–1700 cm−1 range; oxygenation state marker bands are shown. For every RBC, the spectra demonstrate higher (I) and lower (II) oxygenation states.
Figure 5. Characteristic Raman spectra measured in the most oxygenated and in the most deoxygenated areas of an individual RBC: (a) control; (b) RBC treated with TiO2(r); and (c) RBC treated with TiO2(a). Spectra are extended in the 1500–1700 cm−1 range; oxygenation state marker bands are shown. For every RBC, the spectra demonstrate higher (I) and lower (II) oxygenation states.
Materials 14 05920 g005
Figure 6. Deconvolution of characteristic Raman spectra of the most oxygenated Hb (I) and the most deoxygenated Hb (II) in an individual RBC; spectra deconvolution using Lorentzian fitting is shown in green. Inset: the variations of intensities ratio for control RBC and RBC treated with TiO2 are estimated numerically.
Figure 6. Deconvolution of characteristic Raman spectra of the most oxygenated Hb (I) and the most deoxygenated Hb (II) in an individual RBC; spectra deconvolution using Lorentzian fitting is shown in green. Inset: the variations of intensities ratio for control RBC and RBC treated with TiO2 are estimated numerically.
Materials 14 05920 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Perevedentseva, E.; Lin, Y.-C.; Karmenyan, A.; Wu, K.-T.; Lugovtsov, A.; Shirshin, E.; Priezzhev, A.; Cheng, C.-L. Raman Spectroscopic Study of TiO2 Nanoparticles’ Effects on the Hemoglobin State in Individual Red Blood Cells. Materials 2021, 14, 5920. https://doi.org/10.3390/ma14205920

AMA Style

Perevedentseva E, Lin Y-C, Karmenyan A, Wu K-T, Lugovtsov A, Shirshin E, Priezzhev A, Cheng C-L. Raman Spectroscopic Study of TiO2 Nanoparticles’ Effects on the Hemoglobin State in Individual Red Blood Cells. Materials. 2021; 14(20):5920. https://doi.org/10.3390/ma14205920

Chicago/Turabian Style

Perevedentseva, Elena, Yu-Chung Lin, Artashes Karmenyan, Kuan-Ting Wu, Andrei Lugovtsov, Evgeny Shirshin, Alexander Priezzhev, and Chia-Liang Cheng. 2021. "Raman Spectroscopic Study of TiO2 Nanoparticles’ Effects on the Hemoglobin State in Individual Red Blood Cells" Materials 14, no. 20: 5920. https://doi.org/10.3390/ma14205920

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop